You are on page 1of 9

Journal of Mechanical Science and Technology 29 (11) (2015) 4691~4700

www.springerlink.com/content/1738-494x(Print)/1976-3824(Online)

DOI 10.1007/s12206-015-1016-y

An investigation on the supersonic ejectors working with mixture of air and steam
Maziar Shafaee1,*, Mohsen Tavakol1, Rouzbeh Riazi1 and Navid Sharifi2
1
Faculty of New Sciences and Technologies, University of Tehran, 14395-1561, Iran
Departments of Aerospace Engineering, Amirkabir University of Technology, Tehran, 15875-4413, Iran

(Manuscript Received December 13, 2014; Revised May 23, 2015; Accepted June 21, 2015)
----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------

Abstract
This study evaluated the performance of an ejector using two streams of fluids as suction flow. Three motive flow pressures were considered when investigating ejector performance; the suction flow pressure was assumed to be constant. The suction flow consisted of a
mixture of air and steam and the mass fraction of air in this mixture varied from 0 to 1. The ejector performance curves were analyzed for
different mass fractions of air. The results indicate that variation of the mass fraction of air in the suction flow mixture had a significant
effect on ejector performance. At all motive flow pressures, the ejector entertainment ratio increased as the mass fraction of air in the
suction flow increased. The results also show that the sensitivity of ejector performance to variation in the mass fraction of air in the suction flow decreases at higher motive flow pressures. An increase in motive flow pressure caused the transition from supersonic to subsonic flow to occur at higher ejector discharge pressures.
Keywords: Computational fluid dynamics; Entrainment ratio; Steam ejector; Supersonic flow; Working fluid
----------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------

1. Introduction
Ejectors are designed to maximize the suction of secondary
fluid, thoroughly mix the primary and secondary fluids and,
depending on the design requirements, increase the secondary
flow pressure pumped to a region of high pressure. Ejectors
operate with compressible and incompressible fluids and have
applications in the refrigeration, aerospace, water desalination
and petroleum industries [1, 2].
In the refrigeration industry, ejectors have superseded mechanical compressors to compress refrigerant. Ejectors are a
main component of solar cooling systems [3-13] that use renewable energy technology. In desalination systems, ejectors
compress water vapor, which is known as Thermal vapor
compressing (TVC). One application of TVC is for multieffect desalination units. In desalination units a TVC returns
redundant steam to the evaporators, which have higher pressure levels [14].
Keenan et al. [15] initiated theoretical work on ejectors.
Their studies led to the first analytical model of ejectors. The
constant-area mixing model and constant-pressure mixing
model proposed by Keenan provided a framework for ejector
studies. DeFrate and Hoerl [16] improved the constantpressure mixing model by assuming ideal gas consumption
and by taking into account the effect of molecular weight.
*

Corresponding author. Tel.: +98 2161118494, Fax.: +98 2188497324


E-mail address: mshafaee@ut.ac.ir m.tavakol@ut.ac.ir

Recommended by Associate Editor Kyu Hong Kim


KSME & Springer 2015

Munday and Bagster [17] modified the Keenan model to consider chocking and aerodynamic throat theory. Huang et al.
[18] improved the previous 1D analytical model by considering the aerodynamic throat area to give a more accurate prediction of ejector performance. The model provided by Huang
et al. [18] was valid only in the double-choked mode. Chen et
al. [19] improved the Huang model to obtain more accurate
results in the single choked condition.
Analytical models cannot provide sufficient information
about the ejection process because of the complexity of flow
in ejectors. Improvements in computational power and methods have allowed scientists to better understand flow phenomena. Hedges and Hill [20] used Computational fluid dynamics
(CFD) to devise a finite-difference arrangement to simulate
flow inside an ejector. Riffat and Omer [21] studied the performance of a methanol-driven ejector using a commercial
CFD package. Sriveerakul et al. [22] used Fluent software to
simulate the flow inside a steam ejector. They validated their
numerical results using experimental test data. They used
steam as the working fluid by assuming perfect gas and isentropic flow in the ejector.
Hemidi et al. [23] used turbulence models to predict the
ejector entrainment ratio. Simulation results of different turbulence models and experimental data were compared to determine the most accurate models. They found that deviation
from the experimental data was less than 10% for the k-
model; however, this deviation was increased for the k--sst
turbulence model. Bartosiewicz et al. [24] evaluated six turbu-

4692

M. Shafaee et al. / Journal of Mechanical Science and Technology 29 (11) (2015) 4691~4700

Suction Flow

NXP

SuperSonic
flow
Motive
Flow

SubSonic
flow
shock
wave

Mixing
Chamber

OUT FLOW

Constant
Area
Section

Diffuser

Fig. 1. Schematic of a conventional ejector with different flows and geometrical zones.

lence models using CFD simulation of supersonic ejectors.


Axial pressure measurement using a capillary probe showed
that the results of the k--sst model agreed well with the experimental data.
Gagan et al. [25] used Particle image velocimetry (PIV) in a
supersonic gas ejector to find the best turbulence model for
prediction of flow behavior inside an ejector. They found that
turbulence models could predict the flow patterns of vortex
structures and those models like the renormalization group,
the Reynolds stress model and k- could not predict vortex
structures in flow. Moreover, the k- and k--sst models underestimated the scale of the vortex structure when compared
with the experimental results. They found that the best agreement between the CFD results and their PIV data was obtained using the k- standard turbulence model. An optimized
turbulence model depends on ejector operational parameters
and boundary conditions.
Chen et al. [26] stated that one key factor in performance
study of ejectors, used in refrigeration systems, is the working
fluid. The choice of working fluid is essential to the design of
ejectors. They stated some points regarding the thermophysical properties of working fluids. As they mentioned, the
fluid should have high latent heat of vaporization to minimize
the circulation rate per unit of cooling capacity. Furthermore,
the transport properties of working fluid such as its viscosity
and thermal conductivity should be appropriate since these
factors influence the rate of heat transfer. In addition, the effect of molecular mass of the working fluid should be considered for the performance study of ejectors as they stated. In
other words, for a certain system capacity, working fluid with
smaller value of molecular mass requires comparatively larger
ejectors. Also it should be noted that, using the working fluids
with higher molecular mass could lead to an increase in entrainment ratio and ejector efficiency as they mentioned [26].
It can be concluded that working fluid properties, such as its
molecular weight, have significant effects on ejector performance. The effect is more pronounced for ejectors using two or
more fluids. In refrigeration cycles, desalination systems and
other applications, the working fluid could be composed of
fluids with different molecular weights. This could affect ejector performance and make it difficult to predict its behavior.
The present study examined specific operating conditions
on ejector behavior. For this purpose, the effect of the mass

fraction of each species on ejector performance should be


determined; in this regard, it is worth to note that for instance,
in an ejector for desalination systems, a mixture of vapor and
air is sucked into the ejector. The two species have different
molecular weights and affect ejector performance in different
ways. The influence of variation in mass fraction of species in
the suction flow (in terms of variation in mass fraction of air
in the suction flow) on performance of the ejector is studied in
the present work.

2. Operating principles of steam ejectors


2.1 Ejector unit
A typical ejector is composed of a converging-diverging
nozzle (primary nozzle) surrounded by a casing. Fig. 1 shows
the exterior nozzle in three sections; a mixing chamber, a
throat and a diffuser. The primary fluid (motive) coming from
the boiler enters the ejector at high temperature under high
pressure. Since the primary nozzle is a converging-diverging
one, the static pressure of the motive flow converts to velocity
and then accelerates by expanding through the nozzle to supersonic velocity. This supersonic speed causes strong shear
layers between the high velocity flow and the surroundings. A
very low pressure region forms at the Nozzle exit plane (NXP)
and in the mixing chamber. These cause suction of the secondary fluid into the mixing chamber.
The motive flow expands and forms an aerodynamic throat
[18] and the secondary flow accelerates in the foresaid aerodynamic throat until it reaches M = 1 (double-choked flow).
The mixing process is considered to begin at this location. By
the end of the mixing chamber, the two streams are thoroughly mixed. In the ideal case, the high pressure at the diffuser exit creates a shock wave between the end of the constant area and the beginning of the diffuser. This causes major
compression and a sudden drop in flow speed from supersonic
to subsonic. The mixed fluid leaves the subsonic diffuser under middle pressure between the primary and secondary flow
pressures.
2.2 Ejector performance
The Entrainment ratio (ER) and Compression ratio (CR) are
characteristics of an ejector used to evaluate ejector perform-

4693

M. Shafaee et al. / Journal of Mechanical Science and Technology 29 (11) (2015) 4691~4700

ance. The former is non-dimensional mass flow rate and the


latter is non-dimensional pressure. These parameters are defined in Eqs. (1) and (2):
m& suc
m& mot

(1)

CR =

pdis
.
psuc

(2)

In Eq. (1), mmot and msuc are the primary and secondary mass
flow rates, respectively. In Eq. (2), pdis and psuc are the pressure
values at discharge and the suction boundaries, respectively.
Every ejector has a specific performance curve that behaves
differently under different operating conditions. One important parameter is CR. Fig. 2 describes the effect of CR on ER
in the double-choked flow, single-choked flow and the reversed flow regions. If back pressure pdis is less than a certain
critical pressure p*dis, ER remains constant because of choking
of the secondary fluid in the aerodynamic throat (doublechoked flow condition). An increase in back pressure beyond
p*dis prevents the secondary fluid from reaching M = 1 and it
remains un-choked; only the primary stream is choked in the
primary nozzle (single-choked flow condition). A sudden drop
in the secondary mass flow rate can be calculated in this mode.
A further increase in ejector back pressure causes the flow to
return to the ejector and the suction chamber. This operational
condition is known as the reverse flow mode.

3. Model description
3.1 Numerical procedure
Commercial CFD software (Fluent 6.3) was used to study
flow behavior in an ejector. This program employs the finitevolume method for discretization. An implicit density-based
solver solves the nonlinear equations in the steady state condition. For all equations, the convective terms are discretized
using a second-order upwind scheme. The fluid flow inside
the ejector is considered to be axisymmetric [27].
Regarding the employment of boundary condition, it should
be noted that in the present study, since the velocity of the
flow entering and leaving the computational domain is relatively small compared with the supersonic flow condition of
the ejector; there would be no difference between employing
the boundary conditions of stagnation pressure and static pressure. Consequently, for two entering faces of primary nozzle
and secondary inlet stream, the pressure-inlet boundary condition was considered in this work. Also for the outlet of subsonic diffuser of the ejector, the pressure-outlet boundary condition was implemented. It is noteworthy that for all different
operating conditions which have been investigated in this
study, the abovementioned pressure type boundary conditions
were considered.
Since the flow in the ejector is high speed, heat transfer
through the walls is assumed to be negligible.

Single
Choked
mode

Reversed
flow
mode

Entrainment ratio (ER)

ER =

Double Choked
mode

Compression ratio(CR)
Fig. 2. Ejector performance curve.

3.2 Governing equations


The general form of conservation equations for compressible flow in the Cartesian coordinate system are shown in Eqs.
(3)-(5):
r ( r ui )
+
=0
t
xi

(3)

( r ui )

p t ij
( r uiu j ) = +
+
t
x j
xi x j

(4)

(rE)

( r u i E + u i p) =
+
t
xi
p

(k eff
)+
(u i t ij ) .
+
t xi
xi
x j

(5)

Since the working fluid is assumed to comprise two gases in


this study (pure air and water vapor), conservation equations
for the gas species are solved using Eqs. (6) and (7):
N

w = 1
i

(6)

i =1

wi =

mi
mtot

(7)

where N is the total number of gas species.


3.3. Turbulence modeling
Regarding the study of various turbulence models, 5 different turbulence models (i.e., k- realizable, k- RNG, k- standard, k--sst, and k- standard) were investigated in this work.
Fig. 3 shows the variation of Mach number on the centerline
of ejector, using the k- realizable turbulence model, for operating condition with motive flow pressure of 270 kPa considering the both cases of pure steam as the secondary inlet
stream and pure air as the secondary inlet stream. The same

4694

M. Shafaee et al. / Journal of Mechanical Science and Technology 29 (11) (2015) 4691~4700

Table 1. Ejector geometrical parameters.


Ejectors geometry

Dimension (mm)

Throat length (L)

95

Primary nozzle throat

Primary nozzle exit diameter

Mixing chamber inlet diameter (Z)

24

Table 2. The results of grid dependency study for pure steam as ejector
working fluid.
Fig. 3. Variation of Mach number along the centerline of ejector for
cases of pure steam (Black line) and pure air (Red line) as the secondary inlet stream.

calculations were also performed using other turbulence models (i.e., k- RNG, k- standard, k--sst, and k- standard),
however, their results have not been presented here for the
sake of brevity. It should be noted that, among these studied
turbulence models, the results of using turbulence model of k realizable for the case of pure steam were in good agreement
with related experimental results [22].
The case of pure air as the secondary inlet stream was investigated because it was assumed that the study on distribution of the air mass fraction in the flow field and degree of
mixing could become more feasible by increasing the amount
of air in the suction stream. Sharp variations (with high gradient) in the amount of Mach number, on the centerline of ejector, could be assumed as a measure to determine the regions
where strong or weak shocks might occur. It is noteworthy
that the results of k- realizable with pure air show the same
behavior as those of pure steam (see Fig. 3). However, as indicated in Fig. 3, the maximum difference for the amounts of
calculated Mach number between the cases of pure air and
pure steam is occurred in the range of around x~0.12 and
x~0.25.
Based on the obtained results from various turbulence models, the turbulence model of k- realizable was considered as
the suitable method (compared with other turbulence models)
for modeling the effects of variation in mass fraction of the
suction flow in the present study. To calculate flow near the
wall, the logarithmic standard wall function is used. It calculates flow velocity for near-wall nodes using empirical coefficients. The main governing equations of this model are expressed in Eqs. (8) and (9):
m


( r k ) + ( rk ui ) =
m + t
t
x
x j
sk

k
x + sk
j

(8)


( re ) + ( re ui ) =
m + t
t
x
x j
se

e
+ se

x j

(9)

where sk and s are source terms for turbulent kinetic energy


() and turbulent dissipation rate (), respectively, and k and
are the liquid surface tension for and , respectively.

Case

Mesh nodes

Entrainment ratio

18740

0.5273

26700

0.5456

39000

0.5456

Fig. 4. The considered mesh for computational domain.

3.4 Working fluid properties


Air and steam are used as working fluids and ideal gas behavior is assumed when relating pressure to density. This may
cause deviation from experimental results; however, the results are still accurate because the effects of this assumption
on the overall characteristic parameters are less than 5% for
motive flow and 2.3% for suction flow [28].
3.5 Ejector geometry and grid generation
The ejector geometry was selected based on the experimental work of Sriveerakul et al. [22]. The specifications listed in
Table 1 were considered to be the geometrical parameters of
the ejector. Gambit 2.4.6 was used to generate the mesh, employing structured quadrilateral elements. In order to increase
the accuracy of numerical analysis, a grid structure with high
resolution was considered for the regions with high gradients
of flow quantities (see Fig. 4).
Regarding the grid dependency study, the results of 5 different turbulence models (i.e., k- realizable, k- RNG, k-
standard, k--sst, and k- standard) were investigated. The
operating condition with motive flow pressure of 270 kPa for
the both cases of pure steam as the secondary inlet stream and
mixture of air and steam as the secondary inlet stream (or
suction flow) with wair = 0.5, was considered in this section.
The results of grid dependency study have been summarized
in Tables 2 and 3 (regarding the both cases of pure steam and

4695

M. Shafaee et al. / Journal of Mechanical Science and Technology 29 (11) (2015) 4691~4700

Table 3. The results of grid dependency study for mixture of air and
steam as ejector working fluids.
Case

Mesh nodes

Entrainment ratio

18740

0.440

26700

0.445

39000

0.445

Fig. 6. Contour of mass fraction distribution for the case of mixture of


air and steam as suction flow and steam as motive flow (wair = 0.5).

ing conditions (saturated temperatures of primary flow). As


can be seen in Table 4, the CFD results for critical back pressure were underestimated compared with experimental results.
Possible reasons for this error may come from, the difficulty
of calibrating the absolute pressure transducers [22].

4. Results and discussion


4.1 Species distribution
Fig. 5. Effect of grid resolution on variation of Mach number along the
centerline of ejector for the case with mixture of air and steam as the
suction stream (wair = 0.5).

the mixture of air and steam respectively) in terms of the influence of grid density on the calculated values of entrainment
ratio using the turbulence model of k- realizable.
As shown in Tables 2 and 3, by increase of the gird resolution from 18740 up to 27600 the amount of computed entrainment ratio increases. However, increasing the grid density
beyond the amount of 27600 would not influence the calculated entrainment ratios. Also the effect of grid resolution on
variation of Mach number along the centerline of ejector (for
the case of wair = 0.5) has been studied as shown in Fig. 5.
Here again, the results show that by increase of the grid density beyond a certain value, the results of variation of Mach
number do not change.
3.6 Validation of CFD results
The data from Sriveerakul et al. [22] was used to validate
the CFD results. The ejector used in their study was modeled
under similar operating conditions. The simulation incorporates vapor stream as the motive flow. A comparison of the
results listed in Table 4 shows that the numerical results are in
good agreement with the experimental results under all operat-

After initial validation, the analysis can be extended to the


condition where the secondary inlet stream includes both air
and steam. For this purpose, equations for air and steam flow
should be considered. Fig. 6 shows the distribution of air and
steam in the ejector for the case where steam (saturated water
vapor) is the motive fluid and mixture of air and steam with
equal ratios of mass fractions, are the secondary fluids (air =
0.5). The blue region denotes the distribution of water vapor in
the ejector and the red regions denote the distribution of pure
air and demonstrate how mixing varies the local mass fraction.
The figure shows that mixing of the gas species begins at
the exit plane of the converging-diverging nozzle (NXP); mixture becomes uniform at the end of the constant area. A supersonic core can be observed along the center line of the ejector.
This core is composed of pure water vapor; this supersonic
core gradually breaks up and the water vapor mixes with the
surrounding air. A homogeneous mixture of air and vapor
ultimately leaves the ejector.
4.2 Performance characteristics
Three operating conditions for the pressure of motive flow
were considered in this study. A motive flow pressure of
198.5 kPa was assumed to be the operating condition for
a and motive flow pressures of 270 kPa and 361 kPa were

Table 4. Comparison between results of numerical modeling and the experimental data [22].
Operating conditions
Symbol

Entrainment ratio

Primary flow
Secondary flow
Saturated tempera- saturated temperature
ture (Motive flow)
(Suction flow)

Critical back pressure (mbar)

Experiment

CFD results
(Present study)

Deviation from
experimental results
[22] (%)

0.53

0.5445

-1.88

37

33

Experimental
results [22]

CFD results
(Present study)

120

10

130

10

0.4

0.40251

-2.5

50

45

140

10

0.28

0.30355

3.57

65

60

4696

M. Shafaee et al. / Journal of Mechanical Science and Technology 29 (11) (2015) 4691~4700
0.8
0.7
0.6

0.6

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

w=1
w = 0.67
w = 0.5
w = 0.33
w=0

0.7

ER

ER

0.8

w=1
w = 0.67
w = 0.5
w = 0.33
w=0

Fig. 7. Ejector performance curve for operating condition of , considering different values for mass fraction of air at suction (wair).
0.8

0.5

ER

ER

0.5
0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1
2

CR

Fig. 8. Ejector performance curve for operating condition of , considering different values for mass fraction of air at suction (wair).

assumed to be the operating conditions for and g,


respectively (Table 4). Numerical simulations were carried
out to determine ejector performance under each operating
condition; the outlet pressure was continuously decreased
while the suction and motive pressure remained constant. The
mass fraction of air (wair) was varied from 0 to 1 to investigate
the effect of variation in mass fraction of air in suction flow on
the performance of the ejector. The performance curves for ,
and are shown in Figs. 7-9, respectively.
Figs. 7-9 show that ejector performance changes as the
mass fraction of air changes under specific operating conditions. For all operating conditions, as the mass fraction of air
in the suction flow increased, the ejector ER increased.
The most obvious difference between the performance
curves and conventional performance curves of the ejectors
with a single fluid is the variation of ER caused by variation in
working fluid mass fraction. This is mainly the result of the
difference in density between the working fluids in the suction

a (w=1)
a (w=0)
b (w = 1)
b (w = 0)
g (w=1)
g (w=0)

0.7
0.6

0.8

0.6

Fig. 9. Ejector performance curve for operating condition of , considering different values for mass fraction of air at suction (wair).

w=1
w = 0.67
w = 0.5
w = 0.33
w=0

0.7

4
CR

CR

CR

Fig. 10. The comparison of ejector performance for three different


motive pressures (, and ) considering various mass fractions of air
at suction (wair).

flow of the ejector [29]. Figs. 7-9 have been integrated in Fig.
10 to demonstrate how the ER changes as mass fraction of air
changes at different motive flow pressures.
Fig. 10 shows that as the pressure of the motive fluid increased, the ER values remained constant for an extended
range of CR values, which further signifies that increasing the
pressures increases stability. Fig. 10 also show how ER varies
as the mass fraction of air varies. All performance curves indicate an increase in ER with an increase in the mass fraction of
air; however, an increase in ER upon an increase in the mass
fraction of air in the suction flow differed for different motive
flow pressures (Fig. 10).
The variation in ER as a function of variation in the mass
fraction of air in the suction flow for g is lower than for
a and b. In other words, the sensitivity of ejector performance
to the variation in mass fraction of air in the suction flow decreased at higher motive flow pressures (higher momentum of

M. Shafaee et al. / Journal of Mechanical Science and Technology 29 (11) (2015) 4691~4700

the supersonic jet). This might be a result of the decrease in


the growth rate of the shear layer spreading between the motive flow and suction flow at higher motive flow pressures
with the higher momentum of the supersonic jet.
Generally, entrainment takes place in a shear layer which
acts as an interface between motive flow and suction flow
driven by the shear forces existing between the fluids. The
fluid entrained in this shear layer could also cause the layer to
spread and develop. On the other hand, subsonic shear layers
have the highest growth rates and the highest entrainment
from the motion of large scale structures within them [31, 32].
In supersonic jets, additional turbulent structures have been
proposed to assist entrainment [33]. It has been suggested [33]
that these structures play an important role in the entrainment
of fluid into the spreading supersonic jets. If so, the lower
growth rate of development of the shear layer at higher motive
flow pressures with higher momentum of the supersonic jet
may decrease the entrainment of the suction flow into the
mainstream and lower the effect of variation in the mass fraction of air in the suction flow on ER.
Fig. 10 shows that an increase in motive flow pressure to
the break point of the performance curve occurs for higher CR,
so the ejector becomes stable for a wider range of CR.
Note that, for operating condition , the mass flow through
the primary nozzle and the momentum of flow is higher than
for and . This higher momentum and mass flow rate for
increases momentum of the jet core in the ejector and results
in movement of the shock to a position beyond those for and
downstream of the ejector [34]. As a consequence, the ejector will operate in the double-choked mode at a higher discharge pressure than for and . The same is true for the
higher discharge pressure of compared with that of , as
illustrated in Fig. 10.

5. Conclusions
The performance of an ejector using two gases as working
flows and the effect of variation of the mass fraction of air in
the suction flow was investigated. To the knowledge of the
authors, little study has been done on the influence of variation
in mass fraction of air in suction flow on ER of the ejector.
With the aid of numerical simulation, the conditions for two
streams of air and steam were analyzed assuming a perfect gas.
The initial numerical analysis for the ejector employed pure
steam as the working fluid. After the validation of the initial
case with experimental data from related studies [22], the effect of different mixtures of air and steam for the suction flow
were investigated. Ejector performance was analyzed for all
feasible values for the mass fraction of air from 0 to 1.
The results indicate that variation of the mass fraction of air
has a significant effect on ER. Increasing the mass fraction of
air in the suction flow increases the ER. It is also suggested
that higher pressure in the motive flow decreases the effect of
the variation in the mass fraction of suction flow on the ejector
ER. The sensitivity of ejector performance in terms of the

4697

amount of entrained fluid from the suction flow decreased as


the pressure of motive flow increased; an increase in the saturated pressure of the primary flow resulted in higher momentum of the supersonic jets. This could be the result of the decrease in the growth rate of the shear layer spreading between
the motive flow and suction flow where there is higher motive
flow pressure. This lower growth rate of the developing shear
layer may decrease the entrainment of the suction flow into
the mainstream and decrease the contribution of variation in
the mass fraction of air in the suction flow on the ER. The
effect of variations in the pressure of the suction flow and
mass fraction of air in the motive flow (if a mixture of air and
steam is assumed in the motive flow) on the performance of
the ejector is a subject of future investigation.

Nomenclature-----------------------------------------------------------------------CR
dis
E

eff
ER
K

M
mi
m&

t
mot
NXP
P

R
S
Suc
T
t
ij
tot
U

*
i, j,k

: Compression ratio
: Discharge flow
: Total energy
: Turbulent dissipation rate
: Effective
: Entrainment ratio
: Thermal conductivity
: Turbulent kinetic energy
: Mach number
: Molar mass
: Mass flow rate
: Dynamic viscosity
: Turbulent viscosity
: Motive flow
: Nozzle exit position
: Static pressure
: Density
: Refrigerant
: Source term
: Suction flow
: Temperature
: Time
: Stress tensor
: Total
: Velocity
: Mass fraction of ith species in the mixture
: Liquid surface tension
: Critical condition
: Space components

References
[1] D. W. Sun and I. W. Eames, Recent developments in the
design theories and applications of ejectors- a review, J.
Inst. Energ., 68 (1995) 65-79.
[2] Y. K. Kim, D. Y. Lee, H. D. Kim, J. Hahn and K. C. Kim,
An experimental and numerical study on hydrodynamic
characteristicsof horizontal annular type water-air ejector,

4698

M. Shafaee et al. / Journal of Mechanical Science and Technology 29 (11) (2015) 4691~4700

J. Mech. Sci. Technol., 26 (9) (2012) 2773-2781.


[3] Z. Aidoun and M. Ouzzane, The effect of operating conditions on the performance of a supersonic ejector for refrigeration, Int. J. Refrig., 27 (2004) 974-984.
[4] I. I. W. Eames, S. Aphornratana and H. Haider, A theoretical and experimental study of a small-scale steam jet refrigerator, Int. J. Refrig., 18 (1995) 378-386.
[5] S. Aphornratana and I. W. Eames, A small capacity steamejector refrigerator: experimental investigation of a system
using ejector with movable primary nozzle, Int. J. Refrig.,
20 (1997) 352-358.
[6] R. H. Yen, B. J. Huang, C. Y. Chen, T. Y. Shiu, C. W.
Cheng and S. S. Chen, Performance optimization for a
variable throat ejector in a solar refrigeration system, Int. J.
Refrig., 36 (2013) 1512-1520.
[7] K. Chunnanond, Ejectors: applications in refrigeration
technology, Renew. Sust. Energ. Rev., 8 (2004) 129-155.
[8] Y. Allouchea, C. Bouden and S. Varga, A CFD analysis of
the flow structure inside a steam ejector to identify the
suitable experimental operating conditions for a solardriven refrigeration system, Int. J. Refrig., 39 (2013) 1-10.
[9] K. B. Nguyen, S. H. Yoon and J. H. Choi, Effect of working-fluid filling ratio and cooling-water flow rate on the
performance of solar collector with closed-loop oscillating
heat pipe, J. Mech. Sci. Technol., 26 (1) (2012) 251-258.
[10] Y. Allouchea, C. Boudena and S. Riffat, A solar-driven
ejector refrigeration system for Mediterranean climate:
Experience improvement and new results performed, Energ. Procedia., 18 (2012) 1115-1124.
[11] L. Boumaraf and A. Lallemand, Modeling of an ejector
refrigerating system operating in dimensioning and offdimensioning conditions with the working fluids R142b
and R600a, Appl. Therm. Eng., 29 (2009) 265-274.
[12] E. Rusly, L. Aye, W. Charters and A. Ooi, CFD analysis
of ejector in a combined ejector cooling system, Int. J. Refrig., 28 (2005) 1092-1101.
[13] R. Yapici and F. Akkurt, Experimental investigation on
ejector cooling system performance at low generator temperatures and a preliminary study on solar energy, J. Mech.
Sci. Technol., 26 (11) (2012) 3653-3659.
[14] H. Chung, S. Wibowo, B. Fajar, Y. Shin and H. Jeong,
Study on low pressure evaporation of fresh water generation system model, J. Mech. Sci. Technol., 26 (2) (2012)
421-426.
[15] J. H. Keenan and E. P. Neumann, A simple air ejector, J
Appl. Mech-T. Asme., 64 (1942) A75-A81.
[16] L. A. DeFrate and A. E. Hoerl, Optimum design of ejectors using digital computers, Chem. Eng. Prog. Symp., 55
(1959) 43-51.
[17] J. T. Munday and D. F. Bagster, A new ejector theory
applied to steam jet refrigeration, Ind. Eng. Chem. Proc.
D.D., 16 (1977) 442-449.
[18] B. J. Huang, J. M. Chang, C. P. Wang and V. A. Patrenko,
A 1-D analysis of ejector performance, Int. J. Refrig., 22
(1999) 354-364.

[19] W. Chen, M. Liu, D. T. Chong, J. Yan, A. B. Little and Y.


Bartosiewicz, A 1D model to predict ejector performance
at critical and sub-critical operational regimes, Int. J. Refrig., 36 (2013) 1750-1761.
[20] K. R. Hedges and P. G. Hill, Compressible flow ejectors
Part I Development of a finite-difference flow model, J.
Fluid Eng.-T ASME, 96 (1974) 272-281.
[21] S. B. Riffat and S. A. Omer, CFD modeling and experimental investigation of an ejector refrigeration system using methanol as the working fluid, Int. J. Energ. Resv., 25
(2001) 115-128.
[22] T. Sriveerakul, S. Aphornratana and K. Chunnanond,
Performance prediction of steam ejector using computational fluid dynamics: part 1, Validation of the CFD results,
Int. J. Therm. Sci., 46 (2007) 812-822.
[23] A. Hemidi, F. Henry, S. Leclaire, J. M. Seynhaeve and Y.
Bartosiewicz, CFD analysis of a supersonic air ejector.
Part I: experimental validation of single phaseand twophase operation, Appl. Therm. Eng., 29 (2009) 1523-1531.
[24] Y. Bartosiewicz, Z. Aidoun, P. Desevaux and Y.
Mercadier, Numerical and experimental investigations on
supersonic ejectors, Int. J. Heat. Fluid. Fl., 26 (2005) 5670.
[25] J. Gagan, K. Smierciew, D. Butrymowicz and J. Karwacki, Comparative study of turbulence models in application to gas ejectors, Int. J. Therm. Sci., 78 (2014) 9-15.
[26] X. Chen, S. Omer, M. Worall and S. Riffat, Recent developments in ejector refrigeration technologies, Renew.
Sust. Energ. Rev., 19 (2013) 629-651.
[27] K. Pianthong, W. Sheehanam, M. Behnia, T. Sriveerakul
and S. Aphornratana, Investigation and improvement of
ejector refrigeration system using computational fluid dynamics technique, Energ. Convers. Manage., 48 (2007)
2556-2564.
[28] J. M. Cardemil and S. Colle, A general model for evaluation of vapor ejectors performance for application in refrigeration, Energ. Convers. Manage., 64 (2012) 79-86.
[29] C. Li and Y. Z. Li, Investigation of entrainment behavior
and characteristics of gas-liquid ejectors based on CFD
simulation, Chem. Eng. Sci., 66 (2011) 405-416.
[30] V. Dvorak and P. Safarik, Supersonic flow structure in
the entrance part of a mixing chamber of 2D, Journal of
Thermal Science, 12 (4) (2003) 343-349.
[31] P. E. Dimotakis, Two-dimensional shear layer entrainment, AIAA Journal, 24 (1986) 1791-1796.
[32] A. Roshko, Structure of turbulent shear flows: A new
look, AIAA Journal, 14 (1976) 1349-1357.
[33] A. Krothpalli, G. Buzyna and L. Lourenco, Streamwise
vortices in an underexpanded axisymmetric jet, Phys. Fluids A: Fluid Dynamics (1989-1993), 8 (1991) 1848-1851.
[34] T. Sriveerakul, S. Aphornratana and K. Chunnanond,
Performance prediction of steam ejector using computational fluid dynamics: Part 2. Flow structure of a steam
ejector influenced by operating pressures and geometries,
Int. J. Therm. Sci., 46 (2007) 823-833.

M. Shafaee et al. / Journal of Mechanical Science and Technology 29 (11) (2015) 4691~4700

Maziar Shafaee received B.S. degree in


Mechanical Engineering from Tabriz
University in 2000. He also received
M.S. and Ph.D. degrees from University
of Tehran in 2002 and 2011, respectively. Dr. Shafaee is currently worked
as an assistant professor, in Faculty of
New Sciences and Technology at University of Tehran. His research interests include numerical and
experimental study of spray and atomization systems.

4699

Mohsen tavakol received B.S. degree in


Mechanical Engineering from Azad
University in 2012. He is a master student in University of Tehran from 2012.
Mr. Tavakol is currently working on his
M.S. Thesis at Faculty of New Sciences
and Technology in University of Tehran.

You might also like